• Non ci sono risultati.

BETHE ANSATZ AND THE SPECTRAL THEORY OF AFFINE LIE ALGEBRA–VALUED CONNECTIONS. THE SIMPLY–LACED CASE

N/A
N/A
Protected

Academic year: 2021

Condividi "BETHE ANSATZ AND THE SPECTRAL THEORY OF AFFINE LIE ALGEBRA–VALUED CONNECTIONS. THE SIMPLY–LACED CASE"

Copied!
29
0
0

Testo completo

(1)

arXiv:1501.07421v4 [math-ph] 16 Jun 2016

BETHE ANSATZ AND THE SPECTRAL THEORY OF AFFINE LIE ALGEBRA–VALUED CONNECTIONS.

THE SIMPLY–LACED CASE

DAVIDE MASOERO, ANDREA RAIMONDO, DANIELE VALERI

Abstract. We study the ODE/IM correspondence for ODE associated to b

g-valued connections, for a simply-laced Lie algebra g. We prove that subdom-inant solutions to the ODE defined in different fundamental representations satisfy a set of quadratic equations called Ψ-system. This allows us to show that the generalized spectral determinants satisfy the Bethe Ansatz equations.

Contents

0. Introduction 1

1. Affine Kac-Moody algebras and their finite dimensional representations 4

2. Asymptotic Analysis 9

3. The Ψ-system 14

4. The Q-system 18

5. g-Airy function 22

Appendix A. Action of Λ in the fundamental representations 24

References 27

0. Introduction

The ODE/IM correspondence, developed since the seminal papers [13, 3], con-cerns the relations between the generalized spectral problems of linear Ordinary Differential Equations (ODE) and two dimensional Integrable Quantum Field The-ories (QFT), or continuous limit of Quantum Integrable Systems. Accordingly, the acronym IM usually refers to Integrable Models (or Integrals of Motion). More re-cently, it was observed that the correspondence can also be interpreted as a relation between Classical and Quantum Integrable Systems [27], or as an instance of the Langlands duality [17].

The original ODE/IM correspondence states that the spectral determinant of certain Schrödinger operators coincides with the vacuum eigenvalue Q(E) of the Q-operators of the quantum KdV equation [4], when E is interpreted as a parameter of the theory. This correspondence has been proved [13, 3] by showing that the spectral determinant and the eigenvalues of the Q operator solve the same functional equation, namely the Bethe Ansatz, and they have the same analytic properties.

The correspondence has soon been generalized to higher eigenvalues of the Quan-tum KdV equation – obtained modifying the potential of the Schrödinger operator [5] – as well as to Conformal Field Theory with extended Wn-symmetries, by

consid-ering ODE naturally associated with the simple Lie algebra An(scalar linear ODEs

of order n+ 1) [10,34]. Within this approach, the original construction corresponds to the Lie algebra A1, and the associated ODE is the Schrödinger equation.

(2)

After these results, it became clear that a more general picture including all simple Lie algebras – as well as massive (i.e. non conformal) Integrable QFT – was missing. While the latter issue was addressed in [20,27] (and more recently in [11,1,30]), the ODE/IM correspondence for simple Lie algebras different from An

has not yet been fully understood.

The first work proposing an extension of the ODE/IM correspondence to clas-sical Lie algebras other than An is [9], where the authors introduce a g-related

scalar ODE as well as the important concept of Ψ-system. The latter is a set of quadratic relations that solutions of the g-related ODE are expected to satisfy in different representations. It was then recognized in [17] that the g-related scalar differential operators studied in [5,9] can be regarded as affine opers associated to the Langlands dual of the affine Lie algebra bg. This discovery gave the theory a solid algebraic footing – based on Kac-Moody algebras and Drinfeld-Sokolov Lax operators – which was further investigated in [32] and that we hereby follow.

In the present paper we establish the ODE/IM correspondence for a simply laced Lie algebra g, in which case the affine Lie algebrabg coincides with its Langlands dual [17]. More concretely, we prove that for any simple Lie algebra g of ADE type there exists a family of spectral problems, which are encoded by entire functions Q(i)(E),

i = 1, . . . , n = rank g, satisfying the ADE-Bethe Ansatz equations [31,35,2]:

n Y j=1 ΩβjCij Q(j)Cij2 E∗ Q(j)−Cij2 E∗ = −1 , (0.1)

for every E∗ ∈ C such that Q(i)(E) = 0. In equation (0.1), C = (C

ij)ni,j=1 is

the Cartan matrix of the algebra g, while the phase Ω and the numbers βj are the

free-parameters of the equation. More precisely, following [17] we introduce the bg-valued connection

L(x, E) = ∂x+

x+ e + p(x, E)e0, (0.2)

where ℓ ∈ h is a generic element of the Cartan subalgebra h of g, e0, e1, . . . , enare the

positive Chevalley generators of the affine Kac-Moody algebrabg and e =Pni=1ei.

In addition, the potential1

is p(x, E) = xM h∨

− E, where M > 0 and h∨is the dual

Coxeter number of g.

After [32], for every fundamental representation 2

V(i) of bg we consider the

differential equation

L(x, E)Ψ(x, E) = 0 . (0.3)

The above equation has two singular points, namely a regular singularity in x = 0 and an irregular singularity in x = ∞, and a natural connection problem arises when one tries to understand the behavior of the solutions globally. We show that for every representation V(i) there exists a unique solution Ψ(i)(x, E) to equation

(0.3) which is subdominant for x → +∞ (by subdominant we mean the solution that goes to zero the most rapidly). Then, we define the generalized spectral determinant Q(i)(E; ℓ) as the coefficient of the most singular (yet algebraic) term

of the expansion of Ψ(i)(x, E) around x = 0. Finally, we prove that for generic ℓ ∈ h

the generalized spectral determinants Q(i)(E; ℓ) are entire functions and satisfy the

Bethe Ansatz equation (0.1), also known as Q-system.

The paper is organized as follows. In Section 1 we review some basic facts about the theory of finite dimensional Lie algebras, affine Kac-Moody algebras and

1

We stick to a potential of this form for simplicity. However all proofs work with minor modifications for a more general potential discussed in [5,17].

2

Actually V(i)is an evaluation representation of the i-th fundamental representation of g.

(3)

their finite dimensional representations. Furthermore, we introduce the relevant representations that we will consider throughout the paper.

Section 2 is devoted to the asymptotic analysis of equation (0.3). The main result is provided by Theorem2.4, from which the existence of the fundamental (subdom-inant) solution Ψ(i)(x, E) follows. The asymptotic properties of the solution turn

out to depend on the spectrum of Λ = e0+ e ∈bg, as observed in [32].

The main result of Section 3 is Theorem 3.6, which establishes the following quadratic relation among the functions Ψ(i)(x, E), known as Ψ-system, that was

conjectured in [9]: mi Ψ(i)1 2 (x, E) ∧ Ψ(i)1 2 (x, E)= ⊗j∈IΨ(j)(x, E)⊗Bij. (0.4) Here, Ψ(i)±1 2

(x, E) is the twisting of the solution Ψ(i)(x, E) defined in (2.8), while

mi is the morphism ofbg-modules defined by (1.11), and B = (Bij)ni,j=1 denotes

the incidence matrix of g. In order to prove Theorem 3.6, we first establish in Proposition3.4 some important properties of the eigenvalues of Λ on the relevant representations previously introduced. More precisely, we show that in each repre-sentation V(i) there exists a maximal positive eigenvalue λ(i) (see Definition2.3),

and that the following remarkable identity holds:  e−hπi∨ + e πi h∨  λ(i) = n X j=1 Bijλ(j), i = 1, . . . , n = rank g . (0.5)

Equation (0.5) shows that the vector (λ(1), . . . , λ(n)) is the Perron-Frobenius

eigen-vector of the incidence matrix B, and this implies that the λ(i)’s are (proportional

to) the masses of the Affine Toda Field Theory with the same underlying Lie algebra g[18].

In Section4 we derive the Bethe Ansatz equations (0.1). To this aim, we study the local behavior of equations (0.3) close to the Fuchsian singularity x = 0, and we define the generalized spectral determinants Q(i)(E; ℓ) and eQ(i)(E; ℓ) using some

properties of the Weyl group of g. In addition, using the Ψ-system (0.4) we prove Theorem 4.3, which gives a set of quadratic relations among the generalized spec-tral determinants (equation (4.6)). Evaluating these relations at the zeros of the functions Q(i)(E; ℓ), we obtain the Bethe Ansatz equations (0.1).

Finally, in Section 5, we briefly study an integral representation of the sub-dominant solution of equation (0.3) with a linear potential p(x, E) = x, and we compute the spectral determinants Q(i)(E; ℓ). Some explicit computations of the

eigenvectors of Λ are provided in Appendix A.

In this work we do not study the asymptotic behavior of the Q(i)(E)’s, for E ≪ 0,

that was conjectured in [9], neither we address the problem of solving the Bethe Ansatz equations (0.1) – for a given asymptotic behavior of the functions Q(i)(E)’s

– via the Destri-de Vega integral equation [8]. These problems are of a rather different mathematical nature with respect to the ones treated in this work, and they will be considered in a separate paper.

Note added in press. We proved the ODE/IM correspondence for non simply– laced Lie algebras in [29].

(4)

Acknowledgments. The authors were partially supported by the INdAM-GNFM “Progetto Giovani 2014”. D. M. is supported by the FCT scholarship, number SFRH/BPD/75908/2011. D.V. is supported by an NSFC “Research Fund for In-ternational Young Scientists” grant. Part of the work was completed during the visit of A.R. and D.V. to the Grupo de Física Matemática of the Universidade de Lisboa, and during the participation of the authors to the intensive research program “Perspectives in Lie Theory” at the Centro di Ricerca Matematica E. De Giorgi in Pisa. We wish to thank these institutions for their kind hospitality. D.M thanks the Dipartimento di Fisica of the Università degli Studi di Torino for the kind hospitality. A. R. thanks the Dipartimento di Matematica of the Università degli Studi di Genova for the kind hospitality. We also wish to thank Vladimir Bazhanov, Alberto De Sole, Edward Frenkel, Victor Kac and Roberto Tateo for useful discussions.

While we were writing the present paper, we became aware of the work [12] (see [30] for a more detailed exposition) that was being written by Patrick Dorey, Ste-fano Negro and Roberto Tateo. Their paper focuses on ODE /IM correspondence for massive models related to affine Lie algebras and deals with similar problems in representation theory. We thank Patrick, Stefano and Roberto for sharing a preliminary version of their paper.

1. Affine Kac-Moody algebras and their finite dimensional representations

In this section we review some basic facts about simple Lie algebras, affine Kac-Moody algebras and their finite dimensional representations which will be used throughout the paper. The discussion is restricted to simple Lie algebras of ADE type, but most of the results hold also in the non-simply laced case; we refer to [19,24] for further details. At the end of the section we introduce a class ofbg-valued connections and the associated differential equations.

1.1. Simple lie algebras and fundamental representations. Let g be a simple finite dimensional Lie algebra of ADE type, and let n be its rank. The Dynkin diagrams associated to these algebras are given in Table 1. Let C = (Cij)i,j∈I

be the Cartan matrix of g, and let us denote by B = 21n− C the corresponding incidence matrix. Since g is simply-laced, it follows that Bij = 1 if the nodes i, j

of the Dynkin diagram of g are linked by an edge, and Bij= 0 otherwise.

An α1 α2 . . . αn−1 αn E6 α1 α2 α3 α5 α6 α4 Dn α1 α2 . . . αn−2 αn αn−1 E7 α1 α2 α3 α4 α6 α7 α5 E8 α1 α2 α3 α4 α5 α7 α8 α6

Table 1: Dynkin diagrams of simple Lie algebras of ADE type.

(5)

Let {fi, hi, ei | i ∈ I = {1, . . . , n}} ⊂ g be the set of Chevalley generators of g.

They satisfy the following relations (i, j ∈ I):

[hi, hj] = 0 , [hi, ej] = Cijej, [hi, fj] = −Cijfj, [ei, fj] = δijhi, (1.1)

together with the Serre’s identities. Recall that

h=M

i∈I

Chi⊂ g

is a Cartan subalgebra of g with corresponding Cartan decomposition

g= h ⊕ M

α∈R

!

. (1.2)

As usual, the gαare the eigenspaces of the adjoint representation and R ⊂ h∗is the

set of roots. For every i ∈ I, let αi ∈ h∗be defined by αi(hj) = Cij, for every j ∈ I.

Then gαi = Ceαi. Hence, ∆ = {αi | i ∈ I} ⊂ R. The elements α1, . . . , αn are

called the simple roots of g. They can be associated to each vertex of the Dynkin diagram as in Table1. We denote by

P = {λ ∈ h∗| λ(hi) ∈ Z, ∀i ∈ I} ⊂ h∗

the set of weights of g and by

P+ = {λ ∈ P | λ(hi) ≥ 0, ∀i ∈ I} ⊂ P

the subset of dominant weights of g. We recall that we can define a partial ordering on P as follows: for λ, µ ∈ P , we say that λ < µ if λ − µ is a sum of positive roots. For every λ ∈ h∗ there exists a unique irreducible representation L(λ) of g such

that there is v ∈ L(λ) \ {0} satisfying

hiv = λ(hi)v , eiv = 0 , for every i ∈ I .

L(λ) is called the highest weight representation of weight λ, and v is the highest weight vector of the representation. We have that L(λ) is finite dimensional if and only if λ ∈ P+, and conversely, any irreducible finite dimensional representation of

gis of the form L(λ) for some λ ∈ P+. For λ ∈ P+, the representation L(λ) can be

decomposed in the direct sum of its (finite dimensional) weight spaces L(λ)µ, with

µ ∈ P , and we have that L(λ)µ6= 0 if and only if µ < λ. We denote by Pλ the set

of weights appearing in the weight space decomposition of L(λ); the multiplicity of µ in the representation L(λ) is defined as the dimension of the weight space L(λ)µ.

Recall that the fundamental weights of g are those elements ωi∈ P+, i ∈ I,

satis-fying

ωi(hj) = δij, for every j ∈ I . (1.3)

The corresponding highest weight representations L(ωi), i ∈ I, are known as

funda-mental representations of g, and for every i ∈ I we denote by vi∈ L(ωi) the highest

weight vector of the representation L(ωi). Since the simple roots and the

funda-mental weights of g are related via the Cartan matrix (which is non-degenerate) by the relation

ωi=

X

j∈I

(C−1)jiαj, i ∈ I ,

then we can associate to the i−th vertex of the Dynkin diagram of g the corre-sponding fundamental representation L(ωi) of g.

(6)

1.2. The representations L(ηi). Let us consider the dominant weights

ηi=

X

j∈I

Bijωj, i ∈ I ,

where B = (Bij)i,j∈I is the incidence matrix defined in Section 1.1, and let L(ηi)

be the corresponding irreducible finite dimensional representations. In addition, we consider the tensor product representations

O

j∈I

L(ωj)⊗Bij, i ∈ I .

We now show that we can find a copy of the irreducible representation L(ηi) inside

the representations V2L(ωi) and Nj∈IL(ωj)⊗Bij. This will be used in Section 3

to construct the so-called ψ-system.

Lemma 1.1. The following facts hold in the representationV2L(ωi).

(a) For every j ∈ I, we have that

ej(fivi∧ vi) = 0 .

(b) For every h ∈ h, we have that

h(fivi∧ vi) = ηi(h)(fivi∧ vi) .

(c) Any other weight of the representationV2L(ωi) is smaller than ηi.

Proof. Since vi is a highest weight vector for L(ωi) we have that ejvi = 0 for every

j ∈ I. Moreover, using the relations [ej, fi] = δijhi together with equation (1.3),

we have

ejfivi= fiejvi+ δijhivi= δijvi,

and therefore ej(fivi ∧ vi) = δijvi ∧ vi = 0, proving part (a). By linearity, it

suffices to show part (b) for h = hi, i ∈ I. Recall that, by equation (1.1), we have

[hi, fj] = −Cijfj, for every i, j ∈ I. Then,

hjfivi= (fihj+ [hj, fi]) vi = (δij− Cji)fivi.

Hence, hj(fivi∧ vi) = (2δij − Cji)(fivi∧ vi) = ηi(hj)(fivi∧ vi). Finally, let w 6=

ωi, ηi− ωi be a weight appearing in L(ωi). It follows from representation theory of

simple Lie algebras that

ω < ηi− ωi= ωi−

X

j∈I

Cjiωj< ωi.

Therefore, by part (b), the maximal weight of V2L(ωi) is wi+ ηi− ωi = ηi, thus

proving part (c). 

As a consequence of the above Lemma it follows that fivi∧vi∈V2L(ωi) is a highest

weight vector of weight ηi. Since g is simple, the subrepresentation of V2L(ωi)

generated by the highest weight vector fivi ∧ vi is irreducible. Therefore, it is

isomorphic to L(ηi). By the complete reducibility ofV2L(ωi) we can decompose

it as follows:

2

^

L(ωi) = L(ηi) ⊕ U ,

where U is the direct sum of all the irreducible representations different from L(ηi).

By an abuse of notation we denote with the same symbol the representation L(ηi)

and its copy in V2L(ωi). For every i ∈ I, let us denote by wi = ⊗j∈Ivj⊗Bij. The

following analogue of Lemma1.1in the case of the representationNj∈IL(ωj)⊗Bij

holds true.

(7)

(a) For every j ∈ I, we have that

ejwi= 0 .

(b) For every h ∈ h, we have that

hwi= ηi(h)wi.

(c) Any other weight of the representationNj∈IL(ωj)⊗Bij is smaller than ηi.

Proof. Same as the proof of Lemma 1.1. 

As for the previous discussion, by Lemma1.2it follows that wi∈Nj∈IL(ωj)⊗Bij

is a highest weight vector which generates an irreducible subrepresentation isomor-phic to L(ηi). Lemmas 1.1 and 1.2, together with the Schur Lemma, imply that

there exists a unique morphism of representations

mi: 2 ^ L(ωi) → O j∈I L(ωj)⊗Bij, (1.4)

such that Ker mi= U and mi(fivi∧ vi) = wi.

1.3. Affine Kac-Moody algebras and finite dimensional representations. Let h∨ be the dual Coxeter number of g. Let us denote by κ the Killing form of g

and let us fix the following non-degenerate symmetric invariant bilinear form on g (a, b ∈ g):

(a|b) = 1

h∨κ(a|b) .

Let g ⊗ C[t, t−1] be the space of Laurent polynomials with coefficients in g. For

a(t) =PNi=−Maiti∈ C[t, t−1], we let

Rest=0a(t)dt = a−1.

The affine Kac-Moody algebrabg is the vector space bg = g ⊗ C[t, t−1] ⊕ Cc endowed

with the following Lie algebra structure (a, b ∈ g, f (t), g(t) ∈ C[t, t−1]):

[a ⊗ f (t), b ⊗ g(t)] = [a, b] ⊗ f (t)g(t) + (a|b) Rest=0(f′(t)g(t)dt) c ,

[c,bg] = 0 . (1.5)

The set of Chevalley generators ofbg is obtained by adding to the Chevalley genera-tors of g some new generagenera-tors f0, h0, e0(for the construction, see for example [24]).

The generator e0, which plays an important role in the paper, can be constructed

as follows. There exists a root −θ ∈ R, known as the lowest root of g, such that −θ − αi6∈ R, for every i ∈ I. Then

e0= a ⊗ t , for some a ∈ g−θ. (1.6)

We now consider an important class of representations of bg. Let V be a finite dimensional representation of g. For ζ ∈ C∗ we define a representation ofbg, which

we denote by V (ζ), as follows: as a vector space we take V (ζ) = V , and the action ofbg is defined by

(a ⊗ f (t))v = f (ζ)(av) , c v = 0 , for a ∈ g , f (t) ∈ C[t, t−1] , v ∈ V . The representation V (ζ) is called an evaluation representation of bg. If V and W are representations of g then any morphism of representations f : V → W can obviously be extended to a morphism of representations f : V (ζ) → W (ζ), which we denote by the same letter by an abuse of notation. Similarly, when referring to weights or weight vectors of the evaluation representation V (ζ), we mean the weights and weight vectors of the representation V with respect to the action of g.

(8)

For a representation V of g and k ∈ C, we denote Vk= V (e2πik) the evaluation

representation of bg corresponding to ζ = e2πik. Clearly, if k ∈ Z then V

k = V0 =

V (1), and if k ∈ 1

2+ Z then Vk = V12 = V (−1). Then, for each vertex i ∈ I of the

Dynkin diagram of g, we associate the following finite dimensional representation ofbg:

V(i) = L(ωi)p(i)

2 , (1.7)

where L(ωi) is the fundamental representation of g with highest weight ωi, and the

function p : I → Z/2Z is defined inductively as follows:

p(1) = 0 , p(i) = p(j) + 1 , for j < i such that Bij> 0. (1.8)

By an abuse of language we call the representations (1.7), for i ∈ I, the finite dimensional fundamental representations of bg. These representations will play a fundamental role in the present paper.

Example 1.3. In type An, V(1) = L(ω1)0is the evaluation representation at ζ = 1

of the standard representation L(ω1) = Cn+1. Moreover,

V(i)= i ^ Vi−1(1) 2 , for k = 1, . . . , n .

Example 1.4. In type Dn, V(1)= L(ω1)0is the evaluation representation at ζ = 1

of the standard representation L(ω1) = C2n. Moreover,

V(i)= i ^ Vi−1(1) 2 , for k = 1, . . . , n − 2 . Finally, V(n−1)= L(ω n−1)n 2 and V (n)= L(ω n)n

2 are the evaluation representations

at ζ = (−1)n of the so-called half-spin representations.

Example 1.5. In type E6, V(1) = L(ω1)0 and V(5) = L(ω5)0 are the evaluation

at ζ = 1 of the two 27-dimensional representations (they are dual to each other). Moreover, V(2)= 2 ^ V1(1) 2 , V(3)= 3 ^ V0(1)∼= 3 ^ V0(5), V(4)= 2 ^ V1(5) 2 . Finally, V(6) = L(ω6)1

2 is the evaluation representation at ζ = −1 of the adjoint

representation.

As a final remark, note that by equation (1.7) we have

2 ^ L(ωi) ! p(i)+1 2 = 2 ^ V1(i) 2 , (1.9)

and if we introduce the evaluation representation M(i)=O

j∈I

V(j)⊗Bij

, ∀i ∈ I , (1.10)

then by equation (1.7), and from the fact that Bij 6= 0 implies p(i) = p(j) + 1, it

follows that the morphism mi defined by equation (1.4) extends to a morphism of

evaluation representations mi: 2 ^ V1(i) 2 → M (i). (1.11)

The construction of the ψ−system will be provided by means of the above extended morphism. Due to Lemmas1.1and1.2, the evaluation representation

W(i) = L(ηi)p(i)+1

2 , i ∈ I , (1.12)

(9)

is a subrepresentation of both V2V1(i)

2 and M

(i), and the copies of W(i) in these

representations are identified by means of the morphism mi.

1.4. bg-valued connections and differential equations. Let {fi, hi, ei | i =

0, . . . , n} ⊂bg be the set of Chevalley generators of bg, and let us denote e =Pni=1ei.

Fix an element ℓ ∈ h. Following [17] (see also [32]), we consider the bg-valued connection

L(x, E) = ∂x+

x+ e + p(x, E)e0, (1.13)

where p(x, E) = xM h∨

− E, with M > 0 and E ∈ C. Since e is a principal nilpotent element, it follows from the Jacobson-Morozov Theorem together with equation (1.6) that there exists an element h ∈ h such that

[h, e] = e , [h, e0] = −(h∨− 1)e0, (1.14)

see for instance [7]. Under the isomorphism between h and h∗ provided by the

Killing form, the element h corresponds to the Weyl vector (half sum of positive roots). Let k ∈ C and introduce the quantities

ω = eh∨ (M+1)2πi , Ω = e2πiMM+1 = ωh ∨

M.

Moreover, let Mk be the automorphism ofbg-valued connections which fixes ∂x, g

and c and sends t → e2πikt. Then from equations (1.5), (1.13) and (1.14) we get

Mk ωkad hL(x, E)



= ωkL(ωkx, ΩkE) . (1.15)

We denote Lk(x, E) = Mk(L(x, E)), for every k ∈ C. Note in particular that

L0(x, E) = L(x, E). Equation (1.15) implies that for a given finite dimensional

representation V of g, the action of Lk(x, E) on the evaluation representation V0

ofbg is the same as the action of L(x, E) on Vk, for every k ∈ C. In addition, let bC

be the universal cover of C∗, suppose ϕ(x, E) : bC→ V

0 is a family – depending on

the parameter E – of solutions of the (system of) ODE

L(x, E)ϕ(x, E) = 0 , (1.16)

and for k ∈ C introduce the function

ϕk(x, E) = ω−khϕ(ωkx, ΩkE) . (1.17)

Since on any evaluation representation of bg we have

ead ab = eabe−a, a, b ∈bg , (1.18) then by equations (1.15) and (1.18), we have that

Lk(x, E)ϕk(x, E) = 0 , (1.19)

namely ϕk(x, E) : bC→ Vk is a solution of (1.16) for the representation Vk.

2. Asymptotic Analysis

Let V be an evaluation representation ofbg. Let us denote by eC= C \ R≤0, the complex plane minus the semi-axis of real non-positive numbers, and let us consider a solution Ψ : eC→ V of L(x, E)Ψ(x) = Ψ′(x) +  ℓ x+ e + p(x, E)e0  Ψ(x) = 0 . (2.1)

Equation (2.1) is a (system of) linear ODEs, with a Fuchsian singularity at x = 0 and an irregular singularity at x = ∞. Our goal is to prove the existence of solutions of equation (2.1) with a prescribed asymptotic behavior at infinity in a Stokes sector of the complex plane. In the present form, however, equation (2.1) falls outside the reach of classical results such as the Levinson theorem [26] or the

(10)

complex Wentzel-Kramers-Brillouin (WKB) method [16], since the most singular term p(x, E)e0 is nilpotent. In order to study the asymptotic behavior of solutions

to (2.1) we use a slight modification of a gauge transform originally introduced in [32], and we then prove the existence of the desired solution for the new ODE by means of a Volterra integral equation.

We now consider the required transformation. First, we note that the function p(x, E)h1∨ has the asymptotic expansion

p(x, E)h1∨ = q(x, E) + O(x−1−δ), where δ = M (h∨(1 + s) − 1) − 1 > 0, s = ⌊M + 1 h∨M ⌋, (2.2) and q(x, E) = xM + s X j=1 cj(E)xM(1−h ∨ j). (2.3)

For every j = 1, . . . , s, the function cj(E) is a monomial of degree j in E.

Lemma 2.1. Let L(x, E) be thebg-valued connection defined in (1.13) and let h ∈ h satisfy relations (1.14). Then, we have the following gauge transformation

q(x, E)ad hL(x, E) = ∂x+ q(x, E)Λ +

ℓ − M h

x + O(x

−1−δ) , (2.4)

where Λ = e0+ e, the function q(x, E) is given by (2.3) and δ by (2.2).

Proof. It follows by a straightforward computation using relations (1.14).  Remark 2.2. The transformation (2.4) differs from the one used in [32] and given by xMad hL(x, E) = ∂x+ xMΛ + ℓ − M h x − e0E xM(h∨−1).

The latter transformation fails to be the correct if M (h∨− 1) ≤ 1. In fact, in this

case the term e0E/xM(h

−1)goes to zero slowly and alters the asymptotic behavior.

We are then left to analyze the equation

Ψ′(x) +  q(x, E) Λ +ℓ − M h x + A(x)  Ψ(x) = 0 , (2.5)

where A(x) = O(x−1−Mh∨

) is a certain matrix-valued function. From the general theory we expect the asymptotic behavior to depend on the primitive of q(x, E), which is known as the action. In the generic case M+1

h∨M ∈ Z/ +, the action is defined

as

S(x, E) = Z x

0

q(y, E)dy , x ∈ eC, (2.6)

where we chose the branch of q(x, E) satisfying q ∼ |x|M for x real. In the case M+1

h∨M ∈ Z+ then formula (2.6) becomes

S(x, E) = s−1 X j=0 Z x 0 cj(E)yM(1−h ∨ j)dy + c s(E) log x , s = M + 1 h∨M .

Existence and uniqueness of solutions to equation (2.1) (or equivalently, of (2.5)) depend on the properties of the element Λ = e+e0in the representation considered.

We therefore introduce the following

Definition 2.3. Let A be an endomorphism of a vector space V . We say that a eigenvalue λ of A is maximal if it is real, its algebraic multiplicity is one, and λ > Re µ for every eigenvalue µ of A.

(11)

We are now in the position to state the main result of this section.

Theorem 2.4. Let V be an evaluation representation of bg, and let the matrix representing Λ ∈ bg in V have a maximal eigenvalue λ. Let ψ ∈ V be the corre-sponding unique (up to a constant) eigenvector. Then, there exists a unique solution Ψ(x, E) : eC→ V to equation (2.1) with the following asymptotic behavior:

Ψ(x, E) = e−λS(x,E)q(x, E)−h ψ + o(1) as x → +∞ . Moreover, the same asymptotic behavior holds in the sector | arg x| < π

2(M+1), that

is, for any δ > 0 it satisfies

Ψ(x, E) = e−λS(x,E)q(x, E)−h ψ + o(1), in the sector | arg x| < π

2(M + 1)− δ . (2.7) The function Ψ(x, E) is an entire function of E.

Remark 2.5. We will prove in Section 3 that the hypothesis of the theorem are satisfied in the ADE case for the representations V(i), V2

V1(i) 2

, M(i) and W(i),

which were introduced in the previous section.

Remark 2.6. In the case M (h∨− 1) > 1, the asymptotic expansion of the solution

Ψ in the above theorem reduces to

Ψ(x) = e−λxM +1M+1q(x, E)−h ψ + o(1), in the sector | arg x| < π

2(M + 1). Before proving Theorem 2.4, we apply it to prove the existence of linearly inde-pendent solutions to equation (2.1). First, note that for any k ∈ R such that |k| < h∨(M+1)2 , the function

Ψk(x, E) = ω−khΨ(ωkx, ΩkE) , x ∈ R+ (2.8)

defines, by analytic continuation, a solution Ψk : eC→ Vk of equation (2.1) in the

representation Vk. Theorem2.4implies the following result.

Corollary 2.7. For any k ∈ R such that |k| < h∨(M+1)2 , on the positive real semi-axis the function Ψk has the asymptotic behaviour

Ψk(x, E) = e−λγ

kS(x,E)

q(x, E)−hγ−kh(ψ + o(1)) , x ≫ 0 , (2.9) where ψ is defined as in Theorem 2.4 and γ = e2πih∨. Moreover, let l ∈ Z

+ be

such that 1 ≤ l ≤ h∨

2 . Then, the functions Ψ1−l2 , . . . , Ψ l−1

2 are linearly independent

solutions to equation (2.1) in the representation Vl+1

2 .

Proof. A simple computation shows that S(ωkx, ΩkE) = γkS(x, E), and similarly

q(ωkx, ΩkE) = ωMq(x, E). Therefore by Theorem2.4,

Ψ(ωkx, ΩkE) = ω−Mhe−γkλS(x,E)q(x, E)−h(ψ + o(1)) , x ≫ 0 . Hence, on the positive real semi axis we have

Ψk(x, E) = e−γ

kλS(x,E)

q(x, E)−hγ−kh(ψ + o(1)) , x ≫ 0 , where we have used the identity ωM+1 = γ. Note that γk, k ∈ {1−l

2 , . . . , l−1

2 },

are pairwise distinct. Therefore, the vectors ψk are linearly independent. This

concludes the proof. 

Remark 2.8. A slight variation of the proof of Theorem2.4presented below allows one to prove that the asymptotic behavior of Ψ holds on a larger sector of the complex plane, consisting of three adjacent Stokes sectors. This refined result is necessary to consider the lateral connection problem, where one considers the linear

(12)

relations among solutions that are subdominant in different Stokes sectors. Such a problem leads naturally to another instance of the ODE/IM correspondence, namely to spectral determinants related to the transfer matrix of integrable systems, see e.g. [28].

2.1. Proof of Theorem2.4. To complete our analysis, we need to further simplify equation (2.5) with a new gauge transformation. Let us denote by eL(x, E) = q(x, E)ad hL(x, E), where L(x, E) and h are as in Lemma2.1.

Lemma 2.9. Let ℓ = Pi∈Iℓihi and h = Pi∈Iaihi, with ℓi, ai ∈ C, and

con-sider the element N = Pi∈I(ℓi− M ai)fi ∈ g . Then we have the following gauge

transformation:

eα(x) ad NL(x, E) = ∂ + q(x, E) Λ + O(xe −1−M) , where α(x) = (x q(x, E))−1 .

Proof. It follows by a straightforward computation using the explicit form of eL(x, E) given in equation (2.4), together with the definitions of N and α(x) and the com-mutation relations (see (1.1))

[fi, e0] = 0 and [fi, ej] = δijhi, for every i, j ∈ I .

 By Lemma 2.9 and equation (1.18), it follows that the ODE (2.5) is transformed into

e

Ψ′(x) +q(x, E) Λ + eA(x)Ψ(x) = 0 ,e (2.10)

where eA(x) = O(x−1−M), and eΨ(x) = G(x)Ψ(x), with G(x) = eα(x)N. Since

G(x) =1+ o(x

−M), then we look for a solution to equation (2.10) with the same

asymptotic behavior (2.7) as the solution Ψ(x) to equation (2.5). Finally, we define Φ(x) = eλS(x,E)Ψ(x). Then, it is easy to show that equation (e 2.10) takes the form

Φ′(x) +q(x, E) ˜Λ + eA(x)Φ(x) = 0 , (2.11) where ˜Λ = Λ − λ1. Note that ˜Λ ψ = 0. The problem is now reduced to prove the existence of a solution to equation (2.11) satisfying, for every ε > 0, the asymptotic condition

Φ(x) = ψ + o(1) , in the sector | arg x| < π

2(M + 1)− ε . (2.12) To proceed with the proof we need to introduce some elements of WKB analysis, for which we refer to [16, Chap. 3]. We fix κ > 0 (possibly depending on E) big enough such that the Stokes sector

Σ = {x ∈ eC| Re S(x, E) ≥ Re S(κ, E)}

has the following two properties: S is a bi-holomorphic map from Σ to the right half-plane {z ∈ C | Re z ≥ 0}, and Σ coincides asymptotically with the sector | arg x| < π

2(M+1). In other words, any ray of argument | arg x| < π

2(M+1) eventually

lies inside Σ, while any ray of argument | arg x| > π

2(M+1) eventually lies in the

complement of Σ. For any ε > 0, we define Dε= S−1{| arg z| ≤ π2 − ε}, and we

introduce the space B of holomorphic bounded functions U : Dε→ V for which the

limit limx→∞U (x) = U (∞) is well-defined. The space B is complete with respect

to the norm kU k∞= supx∈Dε|U (x)|.

(13)

We construct the solution of equation (2.11) with the desired asymptotic behavior (2.12) as the solution of the following integral equation in the Banach space B:

Φ(x) = K[Φ](x) + ψ(i), (2.13)

where the Volterra integral operator K is defined as

K[Φ](x) = Z ∞ x e− ˜Λ S(x,E)−S(y,E)  e A(y)Φ(y)dy .

The integral above is computed along any admissible path c, and we say that a (piecewise differentiable) parametric curve c is admissible ifR0t| ˙c(s)|ds = O(|c(t)|) and Re S(c(t), E) is a non decreasing function. For example, S−1(t+x) is an

admis-sible path [16]. Note that a solution Φ(x) of equation (2.11), with the asymptotic behavior (2.12) belongs to B. Moreover, it satisfies equation (2.10) if and only if it satisfies the Volterra integral equation (2.13). Indeed, provided that K is a continuous operator, we have that limx→∞K[Φ](x) = 0 and

K[Φ]′(x) = − eA(x)Φ(x) − q(x, E) ˜ΛK[Φ](x) = − eA(x)Φ(x) − q(x, E) ˜ΛΦ(x) − ψ(i)

= − eA(x)Φ(x) − q(x, E) ˜ΛΦ(x) ,

where the last equality follows from the fact that ˜Λψ(i) = 0.

In order to prove that the solution of the integral equation (2.13) exists and it is unique we start showing that K is a continuous operator on B and its spectral radius is zero, i.e. limn→∞kKnk

1

n = 0. By hypothesis, all eigenvalues of Λ have

non-positive real part. Since Λ is semi-simple, i.e. diagonalizable, we deduce that there exists a β > 0 such that for any κ ≥ 0 and any ϕ ∈ V ,

|eκ ˜Λϕ| ≤ β|ϕ| . (2.14)

Since | eA(x)| = O(x−1−M), we have that | eA| ∈ L1(c, | ˙c(t)|dt) for every admissible

path c , and the function

ρc(x) = β

Z ∞ x

| eA(c(s))|| ˙c(s)|ds (2.15)

is well-defined and satisfies limx→∞ρc(x) = 0. Moreover, the inequality (2.14)

implies that Z ∞ x e− ˜Λ S(x,E)−S(y,E)  e A(y)U (y)dy ≤ ρc(x)kU k∞,

for every U ∈ B. By (2.14) and (2.15) it follows [16] that K[U ] is path-independent and thus well defined. Furthermore, we have

|K[U ](x)| ≤ kU k∞ρ(x) and kKk = sup |U|∞=1 kK[U ]k∞≤ ¯ρ , where ρ(x) = inf cadmissible ρc(x) , ρ = sup¯ x∈Dε ρ(x) .

Note that ρ(x) is a bounded function such that limx→∞ρ(x) = 0. Hence, we

conclude that K is a well-defined bounded operator on B. By definition, Kn[U ]

can be written in terms of K(x, y) as

(14)

where K(x, y) is the matrix-valued function

K(x, y) = e− ˜Λ S(x,E)−S(y,E) 

e A(y) , and from this it follows that Kn[U ](x) can be estimated as

|Kn[U ](x)| = kU k∞ Z ∞ x . . . Z ∞ yn−1

| eA(y1)| . . . | eA(yn)|dy1. . . dyn ≤ kU k∞ρ n(x)

n! . Thus

kKnk ≤ ρ¯n

n! , (2.16)

and the series

Φ = X

n∈Z+

Kn[ψ]

converges in B. Clearly Φ satisfies the integral equation (2.13), since

K[Φ] =

X

n=1

Kn[ψ] = Φ − ψ .

Moreover, Φ is the unique solution of equation (2.13), for if ˜Φ is another solution then

Kn[Φ − ˜Φ] = Φ − ˜Φ , for every n ∈ Z +.

By equation (2.16), it follows that Φ = ˜Φ.

It remains to prove that Φ is an entire function of E. This follows, by a stan-dard perturbation theory argument, from the fact that eA and S are holomorphic functions of E in Dε. Theorem2.4is proved.

3. The Ψ-system

In this section, for a simple Lie algebra g of ADE type, we prove an algebraic identity known as Ψ-system, which was conjectured first in [9]. Recall from Section

1.3 that we can write

2 ^ V1(i) 2 ∼ = W(i)⊕ U , M(i)=O j∈I V(j)⊗Bij ∼ = W(i)⊕ eU , (3.1)

where the representation W(i)= L(η

i)p(i)+1

2 , i ∈ I, was defined in equation (1.12),

and U and eU are direct sums of all the irreducible representations different from W(i). Due to (1.11), for every i ∈ I we have a morphism of representations

mi: ^ V1(i) 2 → M (i), Ker m i = U .

The Ψ-system is a set of quadratic relations, realized by means of the morphisms mi, among the subdominant solutions to the linear ODE (2.1) defined on the

rep-resentationsVV1(i) 2 and M

(i).

In order to prove the main result of this section, we need to study the maximal eigenvalue (and the corresponding eigenvector) of Λ in the representationsV2V1(i)

2

, M(i) and W(i). To this aim, we recall some further facts from the theory of simple

Lie algebras and simple Lie groups. First, we need the following

Lemma 3.1. Let h ∈ h satisfy relations (1.14), and let us set γ = e2πih∨. Then, for

every k ∈ C, the following formula holds: γkad hΛ = γkM

−k(Λ) . (3.2)

(15)

In particular, if ψ is an eigenvector of Λ in an evaluation representation V , then ψk = γ−khψ is an eigenvector of Λ in the representation Vk, and we have

Λψ = λψ if and only if Λψk= γkλψk. (3.3)

Proof. It follows directly from the commutation relations (1.14).  Now consider the element ¯Λ = Λ|t=1∈ g, which is well known [25] to be a regular

semisimple element of g. Therefore, its centralizer gΛ¯ = {a ∈ g | [¯Λ, a] = 0} is a

Cartan subalgebra of g, which we denote h′. Due to Lemma3.1, it follows that

γad hΛ = γ ¯¯ Λ , (3.4)

where h ∈ h is the element introduced in Section 1 satisfying relations (1.14). Equation (3.4) says that ¯Λ is an eigenvector of γad h, and this in turn implies that

h′ is stable under the action of γad h. Since γad his a Coxeter-Killing automorphism

of g [25], we can regard γad has a Coxeter-Killing automorphism of h. From these

considerations it follows – as proved in [6] – that we can choose a set of Chevalley generators of g, say {f′

i, h′i, e′i | i ∈ I} ⊂ g, such that h′ =

L

i∈ICh′i and that the

identity

21n+ γ

ad h+ γ− ad h= (Bt)2, (3.5)

holds as operators on h′. Here B denotes, as usual, the incidence matrix of g.

Following [18], relation (3.5) can be used to express – in the basis {h′

i, i ∈ I} –

the eigenvectors of γad h in terms of the eigenvectors of B. Now it is well known

that the incidence matrix B is a non-negative irreducible matrix [24]. Therefore, it has a maximal (in the sense of the Perron-Frobenius theory) eigenvalue, which is γ12+ γ−12, and a corresponding eigenvector, say (x1, x2, . . . , xn) ∈ Rn

>0. If we fix

x1= 1, we then have the following relation for every i ∈ I:

X j∈I Bijxj = (γ− 1 2 + γ 1 2)x i, xi> 0 . (3.6)

Following [18], and essentially using (3.5), we finally obtain ¯

Λ =X

i∈I

γp(i)2 xih′

i, (3.7)

where xi, i ∈ I, are defined in (3.6) and the function p(i), i ∈ I, is given in (1.8).

Remark 3.2. Note that equation (3.7) corresponds to a precise normalization of ¯Λ. In fact, it is always possible to find an automorphism of g, under which h is stable, in such a way that ¯Λ = cPi∈Iγ±

p(i)

2 xih′

ifor any c ∈ C∗and any choice of the relative

phase γ±p(i)2 . The choice of the relative phase is however immaterial. Indeed, under

the transformation γp(i)2 → γ− p(i)

2 the right hand side of (3.7) represents a different

element in the Cartan subalgebra h′which is however conjugated to ¯Λ. Hence, their

spectra coincide.

The decomposition (3.7) allows one to compute the spectrum of Λ using the Perron-Frobenius eigenvector of the incidence matrix B. We will also need the following:

Lemma 3.3. Let αj, j ∈ I, be the positive roots of g with respect to the Cartan

subalgebra h′. Then ¯Λ satisfies

αj(γ− p(i) 2 Λ) = x¯ jγ p(j)−p(i) 2 (1 − γ−2p(j)+1) , for all i, j ∈ I , (3.8)

where the xi are defined in (3.6). This implies that

(16)

Proof. Since Bij6= 0 if and only if p(i) 6= p(j), equation (3.6) implies that X j∈I γ−p(j)2 B ijxj= (γ− 1 2 + γ 1 2)γ p(i)−1 2 x i. (3.9)

Equation (3.9) and the decomposition (3.7) imply the thesis.  Using the above lemma, together with equation (3.7), we compute the maximal eigenvalue of Λ in the representations V(i), VV(i), W(i) and M(i).

Proposition 3.4. Let i ∈ I. If g is a simple Lie algebra of ADE type, then the following facts hold.

(a) For the fundamental representation V(i) there exists a unique maximal

eigen-value λ(i) of Λ. The eigenvalues λ(i), i ∈ I, satisfy the following identity:

X j∈I Bijλ(j)= (γ− 1 2 + γ 1 2)λ(i). (3.10)

We denote by ψ(i)∈ V(i) the unique (up to a constant factor) eigenvector of Λ

corresponding to the eigenvalue λ(i) .

(b) For the representation M(i), we have thatP

j∈IBijλ(i)is a maximal eigenvalue

of Λ. Moreover,

ψ(i)⊗ =

O

j∈I

ψ(j)⊗Bij ∈ M(i) is the corresponding eigenvector.

(c) For the representationV2V1(i) 2

, we have that (γ−1

2+γ

1

2)λ(i)is a maximal

eigen-value of Λ. Moreover, γ−12hψ(i)∧ γ 1 2hψ(i)∈ 2 ^ V1(i) 2

is the corresponding eigenvector. (d) The elements γ−1

2hψ(i)∧ γ12hψ(i) ∈V2V1(i)

2 and ψ

(i)

⊗ ∈ M(i) belong to the

sub-representation W(i). Hence, for the representation W(i) there exists a maximal

eigenvalue µ(i) of Λ, given by

µ(i) =X j∈I Bijλ(i)= (γ− 1 2 + γ 1 2)λ(i),

and the relation

mi γ− 1 2hψ(i)∧ γ 1 2hψ(i)= cψ(i) ⊗

holds for some c ∈ C∗.

Proof. Recall that the weights appearing in the representation L(ωi) are of the

following type:

ωi, ωi− αi, ωi− αi− αj− ω ,

where in the latter case the index j is such that Bij 6= 0 and ω is an integral

non-negative linear combination of the positive roots. Clearly for any ˜h ∈ h′ such that

Re αj(˜h) ≥ 0, ∀j ∈ I we have

Re ωi(˜h) ≥ (ωi− αi)(˜h) ≥ (ωi− αi− αj− ω)(˜h) .

Therefore for any such ˜h the eigenvalues with maximal real part is ωi(˜h), followed

by (ωi− αi)(˜h), (ωi− αi− αj)(˜h), and so on.

We use this simple argument to study the maximal eigenvalues of Λ in the representation V(i). By definition of Λ and ¯Λ, and using equation (3.2), the action

of Λ on V(i) coincides with the action of γp(i)2 (−1+ad h)Λ on L(ω¯

(17)

αj(γ−

p(i)

2 Λ) ≥ 0, for every j ∈ I, and α¯ i(γ− p(i)

2 Λ) = x¯ i(1 − γ1−2p(i)). It follows

that in the representation V(i), Λ has the maximal eigenvalue

ωi(γ−

p(i)

2 Λ) = x¯ i.

Therefore, by definition of the xi’s, the λ(i) satisfy equation (3.6).

Moreover we have

ψ(i)= γp(i)2 hv′

i (3.11)

and

fi′ψ(i) = cγ−hψ(i), (3.12)

for some c ∈ C∗. This completes the proof of part (a). Part (b) follows directly from

part (a). To prove part (c), we follow the same lines we used to prove part (a). In fact, the action of Λ on V1(i)

2 coincides with the action of γ p∗ (i)

2 (−1+ad h)Λ on L(ω¯ i),

where p∗(i) = 1 − p(i). By equation (3.8) we have Re α i(γ

p∗ (i)

2 (−1+ad h)Λ) = 0,¯

while Re αj(γ

p∗ (i)

2 (−1+ad h)Λ) > 0 if p(i) 6= p(j). It follows that the two eigenvalues¯

with maximal real part correspond to to the weights ωi and ωi− αi. By equation

(3.8), these eigenvalues are xiγ

1

2 and xiγ− 1

2, and the corresponding eigenvectors

are γ∓1

2hψ(i). Therefore VV(i) has a unique maximal eigenvalue (γ12 + γ−12)xi,

with eigenvector γ−12hψ(i)∧ γ12hψ(i) = cγ (1+p(i)) 2 h(v′ i∧ f ′ iv ′ i) , (3.13)

for some c 6= 0. Here we have used equations (3.11) and (3.12).

Let us prove (d). By equation (3.11) and Lemma1.2, it follows that ψ(i)⊗ ∈ W(i),

where W(i)is seen as a subrepresentation of M(i). Therefore the maximal eigenvalue

of W(i) and M(i) coincide. Finally by equations (3.11), (3.13) and Lemma1.1, we

have that γ−1

2hψ(i)∧ γ12hψ(i) ∈ W(i) and m

i(γ−

1

2hψ(i)∧ γ12hψ(i)) = cψ(i)

⊗ , c ∈ C∗,

thus proving part (d) and concluding the proof. 

Remark 3.5. We note that we can always choose a normalization of the eigenvectors ψ(i)∈ V(i), i ∈ I, such that

mi γ− 1 2hψ(i)∧ γ 1 2hψ(i)= ψ(i) ⊗ , (3.14)

for every i ∈ I. From now on, we will assume that Proposition 3.4(d) holds with the normalization constant c = 1.

By Proposition3.4(a), for any fundamental representation V(i), i ∈ I ofbg, there

exists a maximal eigenvalue λ(i) with eigenvector ψ(i). Therefore, by Theorem2.4

there exists a unique subdominant solution Ψ(i)(x, E) : eC→ V(i) of the differential

equation (2.1) in the representation V(i) with asymptotic behavior

Ψ(i)(x, E) = e−λ(i)S(x,E)q(x, E)−h ψ(i)+ o(1), in the sector | arg x| < π 2(M + 1).

(3.15) Using Proposition3.4we can prove the following theorem establishing the Ψ-system associated to the Lie algebra g.

Theorem 3.6. Let g be a simple Lie algebra of ADE type, and let the solutions Ψ(i)(x, E) : eC → V(i), i ∈ I, have the asymptotic behavior (3.15). Then, the

following identity holds: mi Ψ(i)1

2

(x, E) ∧ Ψ(i)1 2

(18)

Proof. Due to Proposition 3.4(b) and Theorem2.4, the unique subdominant solu-tion to equasolu-tion (2.1) in M(i) is

⊗j∈IΨ(j)(x, E)⊗Bij = e−µ (i) S(x,E)q(x, E)−hψ(i) ⊗ + o(1)  , for x ≫ 0 .

Moreover, by equation (2.8), Corollary2.7and Proposition3.4(c) we have that Ψ(i)1

2

(x, E) ∧ Ψ(i)1 2

(x, E) = e−µ(i)S(x,E)q(x, E)−hψ(i)1 2

∧ ψ(i)1 2

+ o(1), for x ≫ 0 . The proof follows by equation (3.14) and the uniqueness of the subdominant

solu-tion. 

Example 3.7. In type An, equation (3.16) becomes

mi Ψ (i) −1 2 ∧ Ψ (i) 1 2 

= Ψ(i−1)⊗ Ψ(i+1), for i = 1, . . . , n ,

where we set Ψ(0)= Ψ(n+1)= 1.

Example 3.8. In type Dn, equation (3.16) becomes

mi Ψ (i) −1 2 ∧ Ψ (i) 1 2 

= Ψ(i−1)⊗ Ψ(i+1), for i = 1, . . . , n − 3 , (3.17)

mn−2 Ψ(n−2)1 2 ∧ Ψ (n−2) 1 2  = Ψ(n−3)⊗ Ψ(n−1)⊗ Ψ(n), (3.18) mn−1 Ψ(n−1)1 2 ∧ Ψ(n−1)1 2  = Ψ(n−2)= m n Ψ(n)1 2 ∧ Ψ(n)1 2  , (3.19) where we set Ψ(0)= 1.

Remark 3.9. The Ψ-system in Example 3.8 is a complete set of relations among the subdominant solutions. The analogue system of equations obtained in [32, Eqs. (3.19,3.20,3.21)] lacks of the equation (3.18) for the node n − 2 of the Dynkin diagram. Note that equations [32, Eqs. (3.19,3.20)] are equivalent to (3.17) and (3.19) of the present paper. Indeed, it follows from the direct sum decomposition3.1

that mi◦ ι =1

V(n−2), for i = n − 1, n, where ι is the embedding V(n−2)

ι

֒→V2V1(i) 2

(in fact, V(n−2) = W(n−1)= W(n)in this particular case). The further identity [32,

Eq. (3.21)], which can be obtained from (A.4), requires the introduction of another unknown and therefore equations [32, Eqs. (3.19,3.20,3.21)] are an incomplete system of n equations in n + 1 unknowns.

Example 3.10. In type En, n = 6, 7, 8, equation (3.16) becomes

mi Ψ(i)1 2

∧ Ψ(i)1 2



= Ψ(i−1)⊗ Ψ(i+1), for i = 1, . . . , n − 4 ,

mn−3 Ψ(n−3)1 2 ∧ Ψ (n−3) 1 2  = Ψ(n−4)⊗ Ψ(n−2)⊗ Ψ(n−1), mn−2 Ψ(n−2)1 2 ∧ Ψ (n−2) 1 2  = Ψ(n−3), mn−1 Ψ(n−1)1 2 ∧ Ψ(n−1)1 2  = Ψ(n−3)⊗ Ψ(n), mn Ψ(n)1 2 ∧ Ψ (n) 1 2  = Ψ(n−1), where we set Ψ(0)= 1. 4. The Q-system

In this section, for any simple Lie algebra of ADE type, we define the generalized spectral determinants Q(i)(E; ℓ), i ∈ I, and we prove that they satisfy the Bethe

Ansatz equations, also known as Q-system. The spectral determinants are entire functions of the parameter E, and they are defined by the behavior of the functions

(19)

Ψ(i)(x, E) close to the Fuchsian singularity. More precisely, Q(i)(E; ℓ) is the

coeffi-cient of the most singular term of the expansion of Ψ(i)(x, E) in the neighborhood

of x = 0.

In the case of the Lie algebra sl2, the spectral determinant Q(E; ℓ) was originally

introduced in [14,3], while the generalization to sln has been given in [10,34]. An

incomplete construction for Lie algebras of BCD type can be found in [9,32]. The terminology generalized spectral determinants is motivated by the sl2case. Indeed,

for the Lie algebra sl2, the linear ODE (2.1) is equivalent to a Schrödinger equation

with a polynomial potential and a centrifugal term, and the spectral determinant Q(E; ℓ) vanishes at the eigenvalues of the Schrödinger operator.

Before proving the main result of this section, we briefly review some well-known fundamental results from the theory of Fuchsian singularities. Here we follow [22, Chap. III].

Remark 4.1. In this Section we assume that the potential p(x, E) = xM h∨

is ana-lytic in x = 0, namely M h∨∈ Z

+. With this assumption, the point x = 0 is simply

a Fuchsian singularity of the equation. The case of a potential with a branch point at x = 0 can formally be treated without any modification – see [3] for the case A1

– but the mathematical theory is less developed.

4.1. Monodromy about the Fuchsian singularity. For any linear ODE with Fuchsian singularity at x = 0, namely of the type

Ψ′(x) = A

x + B(x) 

Ψ(x) , (4.1)

where A is a constant matrix and B(x) is a regular function, the monodromy operator is the endomorphism on the space of solutions of (4.1) which associates to any solution its analytic continuation around a small Jordan curve encircling x = 0. More concretely, if we fix a matrix solution Y (x) of the linear ODE (4.1) with non-vanishing determinant (i.e. a basis of solutions), then the monodromy matrix M is defined by the relation

Y (e2πix) = Y (x)M .

Here, Y (e2πix) is the customary notation for the analytic continuation of Y (x) .

We are interested in the invariant subspaces of the monodromy matrix M . By the general theory, the algebraic eigenvalues of the monodromy matrix M are of the form e2πia, for any eigenvalue a of A.

We said that an eigenvalue a of A is non-resonant if there exists no other eigen-value of a′ such that a − a∈ Z. In this case the eigenvalue e2πia has multiplicity

one and there exists a solution of equation (4.1) which is an eigenvector of the monodromy matrix M with eigenvalue e2πia. Such a solution has the form

χa(x) = xa χa+ O(x)



(4.2) for any χa eigenvector of A with eigenvalue a. If the eigenvalues a1, . . . , ak are

reso-nant, meaning that ai− aj∈ Z for i, j = 1 . . . , k, then the eigenvalue e2πia1 = · · · =

e2πiak has multiplicity k. The matrix M , in general, is not diagonalizable when

restricted to the k-dimensional invariant subspace corresponding to this eigenvalue. The corresponding solutions do not have the form (4.2) due to the appearance of logarithmic terms.

We are interested in the linear ODE (2.1), which has a Fuchsian singularity with A = −ℓ, ℓ ∈ h. For any representation V(i), i ∈ I, the eigenvalues for the action

of ℓ on V(i) have the form λ(ℓ), where λ ∈ P is a weight of the representation

V(i). Note that for a generic choice of ℓ ∈ h and for distinct weights λ

1and λ2, the

(20)

eigenvalues are resonant if and only if they coincide, and therefore they correspond to a weight of multiplicity bigger than one.

4.2. The dominant term at x = 0. Let ω ∈ P+, and let us denote by

l∗= max λ∈Pω

Re λ(ℓ) .

Let also λ∗ ∈ P be such that l= Re λ(ℓ). If λis a weight of multiplicity one

in L(ω), then – by the discussion in Section 4.1 – the most singular behavior at x = 0 of a solution to equation (2.1) is x−l∗

. In this section, for any irreducible representation L(ω) of g and a generic ℓ ∈ h, we characterize the weight λ∗ ∈ P

maximizing λ(ℓ) ∈ Z. Moreover, we show that it has multiplicity one. These facts will be used to define the generalized spectral determinant Q(i)(E; ℓ).

It is well-known from the general theory of simple Lie algebras (see [19, Appendix D]), that given a generic element ℓ ∈ h, we can decompose the set of the roots R of gin two distinct and complementary sets (of positive and negative roots)

R+ = {α ∈ R | Re α(ℓ) > 0} , R− = {α ∈ R | Re α(l) < 0} .

Consequently, we can associate to this ℓ ∈ h a Weyl Chamber Wℓ, as well as an

element wℓof the Weyl group (it is the element which maps the original Weyl

cham-ber into Wℓ) and a (possibly new) set of simple roots ∆ℓ = {wℓ(αi) | i ∈ I}. The

weight wℓ(ω) is the highest weight of L(ω) with respect to the new set of simple

roots ∆ℓ, and due to the definition of R−ℓ, the action of any negative simple root

wℓ(−αi), i ∈ I, decreases the value of Re wℓ(ω)(ℓ). Therefore, λ∗ = wℓ(ω) is the

unique weight maximizing the function Re λ(ℓ), λ ∈ Pω. This weight has

multiplic-ity one since it lies in the Weyl orbit of ω.

In the case of a fundamental representation L(ωi), i ∈ I, the weight wℓ(ωi− αi)

belongs to Pωi, and it has multiplicity one. Moreover, it is possible to show that all

the weights in Pωidifferent from wℓ(ωi) are obtained from wℓ(ωi−αi) by a repeated

action of the negative simple roots wℓ(−αi). We conclude that wℓ(ωi− αi) is the

unique weight maximizing Re λ(ℓ) on Pωi \ {wℓ(ωi)}. We have thus proved the

following result:

Proposition 4.2. Let ℓ ∈ h be a generic element, and let wℓthe associated element

of the Weyl group. Then the weight wℓ(ω) ∈ Pω, ω ∈ P+, has multiplicity one and

Re wℓ(λ)(ℓ) > Re λ(ℓ) ,

for any weight λ ∈ Pω, λ 6= wℓ(ω). In the case of a fundamental weight ω = ωi,

i ∈ I, the weight wℓ(ωi− αi) ∈ Pωi has multiplicity one and

Re wℓ(ωi)(ℓ) > Re wℓ(ωi− αi)(ℓ) > Re λ(ℓ) ,

for any λ ∈ Pωi, λ 6= wℓ(ωi), wℓ(ωi− αi).

Let χ(i), ϕ(i) ∈ L(ω

i) be weight vectors corresponding to the the weights wℓ(ωi)

and wℓ(ωi− αi) respectively. As done in Lemmas1.1and1.2it is possible to show

that L(ηi) is a subrepresentation ofV2L(ωi) (respectivelyNL(ωi)⊗Bij) and that

χ(i)∧ ϕ(i) ∈ L(η

i) (respectively ⊗j∈Iχ(j)⊗Bij ∈ L(ηi)). Moreover, χ(i)∧ ϕ(i) and

⊗j∈Iχ(j)⊗Bij are highest weight vectors (with respect to ∆ℓ) of the

subrepresen-tation L(ηi). Hence, we can identify these vectors (up to a constant) using the

morphism mi defined in equation (1.4). From now on we fix the normalization of

χ(i), ϕ(i), i ∈ I, in such a way that

(21)

4.3. Decomposition of Ψ(i). From the above discussion, and in particular form

Proposition 4.2 and equation (4.2), we have that for any evaluation representa-tion Vk(i), i ∈ I and k ∈ C, there exist distinguished (and normalized) solutions χ(i)k (x, E) and ϕ(i)k (x, E) of the linear ODE (2.1) such that they have the most singular behavior at x = 0 and they are eigenvectors of the associated monodromy matrix. Explicitly, these solutions have the asymptotic expansion

χ(i)k (x, E) = x−wℓ(ωi)(ℓ) χ(i)+ O(x),

ϕ(i)k (x, E) = x−wℓ(ωi−αi)(ℓ) ϕ(i)+ O(x).

Since the parameter E appears linearly in the ODE (2.1) and the asymptotic be-havior at x = 0 does not depend on E, by the general theory [22] it follows that χ(i)k (x, E; ℓ) and ϕ(i)k (x, E; ℓ) are entire functions of E.

Using equation (1.17) and comparing the asymptotic behaviors we have that ω−khχ(i)k′(ωkx, ΩkE) = ω −kwℓ(ωi)(ℓ+h)χ(i)(x, E) k+k′ ω−khϕ(i)k′(ω kx, ΩkE) = ω−kwℓ(ωi−αi)(ℓ+h)ϕ(i)(x, E) k+k′, (4.4)

for every k, k′ ∈ C. For generic ℓ ∈ h, the eigenvalues −w

ℓ(ωi)(ℓ), and −wℓ(ωi−

αi)(ℓ) are non-resonant and therefore we can unambiguously define two functions

Q(i)(E; ℓ) and eQ(i)(E; ℓ) as follows. We write

Ψ(i)(x, E, ℓ) = Q(i)(E; ℓ)χ(i)(x, E) + eQ(i)(E; ℓ)ϕ(i)(x, E) + v(i)(x, E) , (4.5) where v(i)(x, E) belongs to an invariant subspace of the monodromy matrix

corre-sponding to lower weights. We call Q(i)(E; ℓ) and eQ(i)(E; ℓ) the generalized spectral

determinants of the equation (1.13). Since Ψ(i)(x, E, ℓ), χ(i)(x, E) and ϕ(i)(x, E)

are entire functions of the parameter E, then also Q(i)(E; ℓ) and eQ(i)(E; ℓ) are

entire functions of the parameter E.

Finally, the Ψ-system (3.16) implies some remarkable functional relations among the generalized spectral determinants. Indeed, we have the following result. Theorem 4.3. Let ℓ ∈ h be a generic element. Then, the spectral determinants Q(i)(E; ℓ) and eQ(i)(E; ℓ) are entire functions of E and they satisfy the following

Q eQ-system: Y

j∈I

Q(j)(E; ℓ)Bij = ω−12θiQ(i)(Ω−12E; ℓ) eQ(i)(Ω 1 2E; ℓ)

− ω12θiQ(i)(Ω21E; ℓ) eQ(i)(Ω−12E; ℓ) ,

(4.6)

where θi= wℓ(αi)(ℓ + h).

Proof. We already showed that Q(i)(E; ℓ) and eQ(i)(E; ℓ) are entire functions of the

parameter E. By the definition of Ψ(i)±1

2(x, E; ℓ) given in (2.8) and from equations

(4.5) and (4.4) we have that Ψ(i)±1 2 (x, E; ℓ) = ω±12wℓ(ωi)(ℓ+h)Q(i)(Ω±12E; ℓ)χ(i) 1 2 (x, E) + ω±12wℓ(ωi−αi)(ℓ+h)Qe(i)(Ω±12E; ℓ)ϕ(i)1 2 (x, E) + ˜v (i)(x, E) . (4.7)

Here we used the fact that, since V(i)1 2 = V1(i) 2 , then χ(i)1 2 (x, E) = χ(i)1 2 (x, E) and ϕ(i)1 2 (x, E) = ϕ(i)1 2

(x, E). Note that ˜v(i)(x, E) belongs to an invariant subspace of

the monodromy matrix corresponding to lower weights. Equation (4.6) is then obtained by equating the component in W(i) in the ψ-system (3.16) by means of

equations (4.5), (4.7) and the normalization (4.3). 

(22)

We now derive the Q-system (Bethe Ansatz) associated to the Lie algebras of ADE type. Let us denote by Ei ∈ C any zero of Q(i)(E; ℓ). Evaluating equation

(4.6) at E = Ω±1

2Ei we get the following relations

Y j∈I Q(j)(Ω12E; ℓ)Bij = −ω12θiQ(i)(ΩE i; ℓ) eQ(i)(Ei; ℓ) , Y j∈I Q(j)(Ω−12E; ℓ)Bij = ω−12θiQ(i)(Ω−1E i; ℓ) eQ(i)(Ei; ℓ) . (4.8)

Assuming that for generic ℓ ∈ h the functions Q(i)(E; ℓ) and eQ(i)(E; ℓ) do not have common zeros, and taking the ratio of the two identities in (4.8), we get, for any zero Ei of Q(i)(E; ℓ), the Q-system

n Y j=1 ΩβjCij Q(j)Cij2 E∗ Q(j)−Cij2 E∗ = −1 , (4.9)

where C = (Cij)i,j∈I is the Cartan matrix of the Lie algebra g, and

βj = 1 2M h∨ X i∈I (C−1)ijθi= 1 2M h∨wℓ(ωj)(ℓ + h) , j ∈ I .

Remark 4.4. The construction of the Q eQ system (4.6) works for any affine Kac-Moody algebra. However, formula (4.9) holds only in the ADE case. In order to define the analogue of equation (4.9) for the Cartan matrix of a non-simply laced algebra g, one has to consider, as suggested in [17], a connection with values in the Langlands dual of the affine Lie algebrabg.

4.4. The case ℓ = 0. The case of ℓ = 0 is not generic and thus the Q functions cannot be defined as in Section 4. It is however straightforward to generalize the generic case to ℓ = 0, either directly or as the limit ℓ → 0. Indeed, we can take the limit ℓ → 0 along sequences such that the element of the Weyl group wℓ = w

is fixed. In this way, the limit solutions χ(i)(x, E, ℓ = 0), ϕ(i)(x, E, ℓ = 0) of (2.1)

satisfy the Cauchy problem

χ(i)k (x, E, ℓ = 0) = vw i , ϕ

(i)

k (x, E, ℓ = 0) = fiwvwi . (4.10)

Here vw

i and fiware respectively the highest weight vector of V(i) and the negative

Chevalley generator corresponding to the element w of the Weyl group. Equations (4.10) define uniquely the two solutions. Then the spectral determinants Q(i), ˜Q(i)

are exactly the coefficients of Ψ(i)(x = 0, E) with respect to the basis element

vw

i , fiwvi. Note that this implies a certain freedom in the choice of the spectral

determinants Q(i), ˜Q(i). In fact, the freedom in the choice of Q(i) corresponds to

the elements of the orbit of ωi under the Weyl group action, while the freedom

in the choice of ˜Q(i) corresponds to the elements of the orbit of the ordered pair

(ωi, αi) under the same group action.

For any choice of wℓ= w, the functions Q(i), ˜Q(i)satisfy (4.8) with ℓ = 0, wℓ= w

and the corresponding Bethe Ansatz equation (4.9), provided they do not have common zeros.

5. g-Airy function

We consider in more detail the special case of (2.1) with a linear potential p(x, E) = x and with ℓ = 0, which is particularly interesting because the sub-dominant solutions and the spectral determinants have an integral representation. Since in the sl2case the subdominant solution coincides with the Airy function, we

call the solution obtained in the general case the g-Airy function. The classical Airy

(23)

function models the local behavior of the subdominant solution of the Schrödinger equation with a generic potential close to a turning point [16]; we expect the g-Airy function to play a similar role for equation (2.1).

5.1. Q functions. In the special case of (2.1) with a linear potential p(x, E) = x and with ℓ = 0, it is straightforward to compute the Q functions. By the discussion in Section4.4, in order to define the Q functions we need to chose an element w of the Weyl group and thus the distinguished element vw

i of the basis of V(i). The spectral

Q(i)(E) is then defined as the coefficient with respect to vw

i of Ψ(i)(x = 0, E). Due

to the special form of the potential we have

Ψ(i)(x = 0, E) = Ψ(i)(−E, 0) . and thus, using formula (3.15), we get

Q(i)(E) = e−λ(i) h∨

h∨ +1|E| h∨ +1 h∨ Ci 1 + o(1)  as E → −∞ , (5.1)

where Ci is the coefficient of the vector |E|−hψ(i) with respect to the basis vector

vw

i . It is important to note that the asymptotic behavior of the functions Q(i) is

expressed via the eigenvalues Λ(i) which coincide with the masses of the (classical)

affine Toda field theory, as we proved in Proposition3.4. This is precisely the same behavior predicted for the vacuum eigenvalue Q of the corresponding Conformal Field Theory [31].

We conclude that the ODE/IM correspondence depends on the relation among the element Λ ∈bg, the Perron-Frobenius eigenvalue of the incidence matrix B and the masses of the affine Toda field theory.

Remark 5.1. Note that formula (5.1) does not coincide with the asymptotic behavior conjectured in [9, Eq 2.6] for a general potential. In fact, J. Suzuki communicated us [33] that the latter formula is conjectured to hold only for potentials such that

M+1

M h∨ < 1, in which case the Hadamard’s product [9, Eq 2.7] converges. Clearly in

case of a linear potential M+1 M h∨ =

h∨

+1 h∨ > 1.

Remark 5.2. A connection between generalized Airy equations and integrable sys-tems appears, among other places, in [23], where the the orbit of the KdV topolog-ical tau–function in the Sato Grassmannian is described. We note that an operator of the form (1.13) – specialized to the case of algebras of type slnand for ℓ = −h/h∨,

M = 1/h∨ – also appears in that paper. It would be interesting to obtain a more

precise relation between these results and those of the present paper.

5.2. Integral Representation. Given an evaluation representation ofbg, we look for solutions of the equation

Ψ′(x) + e + xe 0  Ψ(x) = 0 , (5.2) in the form Ψ(x) = Z c e−xsΦ(s)ds , (5.3)

where Φ(s) is an analytic function and c is some path in the complex plane. Dif-ferentiating and integrating by parts we get

Z c e−xs − s + e + e0 d ds  Φ(s)ds + e−xse0Φ(s) c= 0 , (5.4)

where the last term is the evaluation of the integrand at the end points of the path. We are therefore led to the study of the simpler linear equation

− s + e + e0 d

ds 

Φ(s) = 0 , (5.5)

which we solve below for two important examples.

(24)

5.3. An−1 in the representation V(1). In this case equation (5.2) is already

known in the literature as the n-Airy equation [16]. Using the explicit form of Chevalley generators ei, i = 0, . . . , n − 1, which is provided in AppendixA.1, and

denoting by Φi(s) the component of Φ(s) in the standard basis of Cn, we find the

general solution of (5.5) as

Φ1(s) = κ e

sn+1

n+1 , Φ

i(s) = si−1Φ1(s) , i = 2, . . . , n ,

where κ is an arbitrary complex number. If we choose the path c as the one connecting e−n+1πi ∞ with en+1πi ∞, then the integral formula (5.3) is well-behaved,

and the boundary terms in (5.4) vanish. Thus, the sln-Airy function has the integral

representation Ψj(x) = 1 2πi Z e+ πin+1 e−n+1πi sj−1e−xs+sn+1n+1ds , j = 1, . . . , n . (5.6)

In case n = 2, the latter definition coincides with the definition of the standard Airy function. By the method of steepest descent we get the following asymptotic expansion of Ψ for x ≫ 0: Ψj(x) ∼ r 1 2πnx 2j−1−n 2n e− n n+1x n+1 n , j = 1, . . . , n . (5.7)

Due to Theorem2.4, the sln-Airy function coincides with the fundamental solution

Ψ(1) of the linear ODE (5.2). In fact the asymptotic behavior of the Airy function

coincides with the asymptotic behavior of the function Ψ(1) (for this computation

one needs to use the explicit formula for h which is given in equation (A.1) ). The solutions Ψ(1)k , k ∈ Z, are obtained by integrating (5.3) along the contour obtained rotating c by e2kπin+1.

5.4. Dn in the standard representation. The second example is the Dn-Airy

function in the representation V(1). Using the Chevalley generators e

i, i = 0, . . . , n,

as described in Appendix A.2, and denoting Φj the jth-component of Φ in the

standard basis of C2n, we find that the general solution of (5.5) reads

Φ1= κ s− 1 2es2n −1 2n−1 Φj= sj−1Φ1(s) , j ≤ n − 1 , Φn= sn−1 2 Φ1(s) , Φn+1= s n−1Φ 1(s) , Φn+j = sn+j−2Φ1(s) , j ≤ n − 1 , Φ2n=  s2n−2 2 + 1 4s  Φ1(s) .

where κ ∈ C is an arbitrary constant. Let c be the path connecting e−2n−1πi ∞ with

e2n−1πi ∞. Then along c the integral formula (5.3) is well-behaved and it thus defines

a solution of (5.2) that we call Dn-Airy function. Moreover, the boundary terms

in (5.4) vanish. As in the case An, the standard method of steepest descent shows

that such solution is exactly the fundamental solution Ψ(1)to equation (5.2). The

solutions Ψ(1)k , k ∈ Z, are obtained by integrating (5.3) along the contour obtained rotating c by e2n−1kπi .

Appendix A. Action of Λ in the fundamental representations In this appendix we give an explicit description of the maximal eigenvalues of Λ in the fundamental representations and of the corresponding eigenvector.

Riferimenti

Documenti correlati

A similar analysis had previously been performed using INTE- GRAL data (Del Santo et al. 2006), but the improved energy resolution and sensitivity of NuSTAR allows us to

Et c’est bien là l’enjeu de cette hésitation entre “commencement” du mot et “vérité” du discours: la représentation politique de la forme du motif de la quête de

The DSR tool defines the switches to operate, identifying the final restoration schemes of the network, for all possible faults. However, the switching sequences are unknown and the

Si tratta di robot rasaerba automatici, che riescono a gestire il taglio del prato, con una modalità che riduce al minimo la necessità di intervento da parte di

Durante il decorso della malattia, infatti, si assiste ad un “adattamento” dei nefroni alla riduzione del loro numero e questo compenso risulta essere sia di tipo anatomico

La linea di prodotto scelta (e selezionata sul portale online della Guida) è ‘l’Or- ganizzazione dei prodotti turistici o servizi turistici’.. Sveva e Mara si propongono di

Senza ovviamente andare alla deriva pop-etica che questi argomenti inevitabilmente producono in una società iperinformata, iperascoltata, iperrappresentata, la

excreted by the insects through the faeces and this would be a plausible explanation for the rapid decline in toxin concentration detected in YMW when larvae fasted